Use of Geotextiles For Reinforcement and

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Geotextiles and Geomembranes 8 (1989) 217-237

Use of Geotextiles for Reinforcement and Strain Relief in


Asphalt Concrete

Robert L. Lytton
Texas A & M University,TexasTransportationInstitute, CE/TTIBuilding, Room 508,
CollegeStation, Texas77843, USA

A BS TRA CT

This paper describes the three uses of geotextiles in asphaltic concrete


overlays: (a) reinforcing (b) strain relief and (c) undersealing. Reflection
cracking through overlays is caused by both loads and thermal contraction,
and cracks may be retarded or arrested by geotextiles if the proper material
properties are selected and good construction techniques are used.
Prediction of crack growth is made with fracture mechanics and this is the
basis of an overlay design computer program which is described herein.
Fracture properties of asphaltic concrete under fatigue loading and thermal
contraction conditions are presented and the way they are altered by the
addition of geotextiles is illustrated. The design equations for the thickness
of an overlay with a reinforcing grid are also given.
Some cautionary rules of thumb, concerning the appropriate pavement
conditions when geotextiles may be expected to prolong the life of an
overlay and when they may not, are also given.

INTRODUCTION

Many pavements which are considered to be structurally sound after the


construction of an overlay, prematurely exhibit a cracking pattern similar
to that which existed in the old pavement. The cracking in the new overlay
surface is due to the inability of the overlay to withstand movements of the
underlying pavement which may be due to traffic loading causing differen-
tial deflections in the underlying pavement layers, expansion or contrac-
tion of subgrade soils, or expansion or contraction of the pavement itself
217
Geotextiles and Geomembranes 0266-1144/89/$03-50 © 1989 Elsevier SciencePublishers
Ltd, England. Printed in Great Britain
218 Robert L. Lytton

due to changes in temperature. These movements create shear or tensile


stresses in the new overlay, which can become greater than the shear or
tensile strength of the asphaltic concrete, and cause cracks to develop in
the new overlay. The propagation of an existing cracking pattern from the
old pavement into and through the new overlay is known as 'reflection
cracking'.
The occurrence of reflection cracking may take place several years after
overlay construction or after only a few months. This form of cracking,
together with its accompanying effects, is the primary cause of overlay
deterioration. When reflection cracking occurs, the continuity of the
overlay surface is destroyed, the structural strength of the pavement is
decreased, and water is allowed to enter the pavement system, leading to
further deterioration. Thus, the occurrence of reflection cracking pre-
maturely shortens the useful life of asphalt overlays by extending the same
problems which weakened the original pavement into the new overlay.
A variety of new materials has been used to retard reflection cracking
and to reduce the amount of water that enters the sublayers once a crack
has occurred. The principal methods are reinforcing fibers, fabrics, strain-
absorbing m e m b r a n e interlayers and reinforcing grids. The three tech-
niques employ one or more of the following in extending the life of an
overlay: (a) reinforcing; (b) strain relief; and (c) undersealing. Geotextiles
can do all three.

C A U S E S OF R E F L E C T I O N C R A C K I N G

Both traffic loads and temperature changes cause cracks in an old pave-
ment to reflect through the overlay. Every time a load passes over a crack
in the old pavement, three pulses of high stress concentrations occur at the
tip of a crack as it grows up through the overlay, as illustrated in Fig. 1. The
first stress pulse is a m a x i m u m shear stress pulse shown at Point A in Fig. 1.
The second stress pulse is a maximum bending stress pulse shown at Point
B in Fig. 1. The third stress pulse is again a maximum shear stress pulse,
except that it is in the opposite direction to the previous shear stress pulse.
Also, because there is a void beneath the old surface course at this point,
the m a x i m u m shearing stress with the load at Point C is larger than when it
is at Point A. These stress pulses occur in a very short period of time, in the
order of 0-05 s. The stiffness of the asphaltic concrete in the overlay and in
the old pavement at these high loading rates is fairly high.
The change of temperature in an overlaid pavement can also cause a
reflection crack to grow. The thermal stresses in the overlay are due to
temperature changes at the surface as shown at Point A in Fig. 2, and to the
Geotextiles for reinforcement of asphalt concrete 219

TIP OF THE CRACK

OVERLAY

OLD SURFACE COURSE

//,~, &"//,~//,~,
,/•//.•//I VOID

POSITION OF WHEEL LOAD

Fig. 1. Stresses and crack growth in overlays due to traffic.

CRACK GROWTH DUE


TO THERMAL STRESS
IN OVERLAY

TEMPERATURE DECREASE /
OVERLAY

OLD SURFACE COURSE

~ ~ ~/~/.~/,~/,'~//
CRACK GROWTH DUE TO
THERMAL CONTRACTION
AND CURLING OF THE
OLD PAVEMENT SURFACE
LAYER

Fig. 2. Temperature changes and consequent crack growth in overlays.


220 Robert L. Lytton

contraction and curling of the underlying old pavement surface as shown at


Point B in the same figure. It is observed that thermal stresses can cause
cracks to propagate both from the top and the bottom of the overlay. The
contraction and curling of the old pavement surface layer apply a shear
stress along the bottom of the overlay and produce a concentration of
tensile stress at Point B. The change of temperature in a pavement occurs
very slowly, over a period of several hours or even the major part of a day.
The stiffness of the asphaltic concrete in the overlay and in the old
pavement is very low, as much as 1000 to 10 000 times lower than the
modulus that these same materials exhibit under traffic loads.
Every time a load passes and every time the temperature decreases in an
overlaid pavement the reflection crack grows a little more. The best
hope of retarding the growth of such reflection cracks is in the selection of
the material properties of the overlay and of the geotextiles e m b e d d e d in it
so as to reduce, as much as is possible, the growth of these cracks.

TESTING REFLECTION CRACKING PROPERTIES


OF O V E R L A Y S

Because reflection cracking has a variety of contributing causes, including


traffic, temperature and moisture changes in the base course and sub-
grade, and because the application of shear and tensile stresses to the
overlay occur at different rates and temperatures, it is difficult to devise a
single laboratory test that will exactly duplicate the behavior of an overlay
under field conditions. Such a laboratory testing device must be set up to
simulate a single chosen field condition in which the traffic, crack spacing,
temperature change in the underlying cracked pavement, subgrade stiff-
ness, presence or absence of voids beneath the pavement surface, degree
of aggregate interlock across the crack in the old pavement, overlay
thickness, geotextile position, and tack coat application rate are specified.
Because each of these affects the rate at which the reflection crack
propagates through the overlay, the extension of the test results to other
field conditions requires a new test set-up and a separate evaluation. The
evaluation of the relative effectiveness of various geotextiles at different
positions within the overlay must be done on a project-by-project basis,
provided that all of the variables that are noted above can be specified and
simulated in the testing apparatus.
There is an alternative to this approach which makes use of relatively
simple tests, each of which determines the 'fracture properties' of an
overlay due to one of the three major means by which cracks propagate
through the overlay: bending, shear and contraction of the underlying
Geotextiles for reinforcement of asphalt concrete 221

pavement. The 'fracture properties' are material properties of the overlay


and depend upon the properties of the asphaltic concrete mixture, the
geotextile, its position within the overlay, and its tack coat application
rate. The 'fracture properties' due to bending, shear and contraction are
then put into a simple computer program along with data on traffic, daily
temperature change, crack spacing and overlay thickness, to calculate the
number of load applications and temperature cycles that a given overlay
can withstand. The advantage of this approach is that more of the simple
tests can be made for each mode of fracture separately to permit a more
careful investigation of the best geotextile properties and positions within
an overlay and the best tack coat application rate, to reduce as much as
possible the rate of crack growth through an overlay. These tests can show
more clearly the contribution of the geotextile to the retardation of
reflection cracks in each fracture mode separately, and can lead more
directly to rules, guidelines and specification limits on the use and applica-
tion of geotextiles in overlays. This is the alternative that has been adopted
by the Texas Transportaton Institute with the assumption that the fracture
properties due to bending and shear are the same.
This was the approach used in the N C H R P Project 20-7, Task 17, on
'Evaluation of A A S H O Road Test Satellite and Environmental Studies', 1
which was subsequently expanded in the F H W A project on the 'Develop-
ment of Asphalt Concrete Overlay Design Equations'. 2 The computer
program which was developed in the latter project takes the 'fracture
properties' of samples with a geotextile embedded in them, as measured in
the beam fatigue and 'overlay tester' laboratory testing devices, and with
them predicts how long an overlay will last under specified traffic and daily
temperature changes.
From the discussion just concluded, it is apparent that when a geotextile
is placed in a pavement overlay, it must counteract the growth of cracks up
from the old pavement due to traffic and thermal stresses. The cracks in
the old pavement include alligator cracks in the wheel paths, due to traffic
loads, and transverse cracks, due principally to thermal stresses. If it were
possible to keep these cracks from moving, i.e. opening and closing,
almost any overlay would last for a very long time. However, the greater
the crack movement, the shorter the life of the overlay, and the greater is
the need for reinforcing or effective strain relief. The greatest amount of
movement occurs as horizontal openings and shearing movements at the
transverse cracks. These are spaced farther apart than the cracks in the
network which makes up alligator cracking. Because a geotextile helps to
protect the overlay against the horizontal movement of the cracks in the
old pavement, there is a need for more reinforcing or strain relief across a
transverse crack than across a wheel path.
222 Robert L. Lytton

SAMPLE F A I L U R E M O D E S IN T H E B E A M F A T I G U E TEST

The beam fatigue test apparatus is shown in Fig. 3. The geotextile is placed
18 mm to 25 mm from the tensile face of the beam. The zone between the
two loads is a constant moment zone in which there are no shear stresses.
Failure occurs gradually as a 'fracture process zone', in which a network of
microcracks develop ahead of a distinctly observable crack. The process
zone works its way through the beam sample from the tensile face to the
compressive face. When the crack finally forms, it moves quickly all the
way through the beam, causes the beam to fail, and the test is shut off
automatically by a limit switch. The crack in the beam is usually a single
crack in the constant moment zone.

P P

-----Geotextile

2¢m I0 cm I0 cm IOcm

R R
Fig. 3. Load distribution and geotextite location in flexural fatigue test.

The effect of a geotextile on the results of this test is to delay the time for
the beam to reach failure. The nature of the delay depends upon whether
the geotextile reinforces or relieves the strain in the fatigue beam. As will
be more apparent later, reinforcement can occur only if the geotextile has
a modulus that is larger than that of the surrounding asphaltic concrete.
Strain relief will occur if its modulus is smaller.

SAMPLE F A I L U R E M O D E S IN T H E O V E R L A Y TEST

In making tests with the overlay tester, it has been discovered that the
mode of failure reveals the nature of the material that has been placed in
the overlay. Three distinct modes of failure are observed as shown in Fig.
4. The 'failure' of the sample was defined for each mode of failure as
follows:
Mode I - - C r a c k penetrates to the top of the sample.
Geotextiles for reinforcement of asphalt concrete 223

Mode II - - T h e force required to open the crack reaches a m i n i m u m


plateau.
Mode III--Cracks meet in the middle of the sample.
Modes I and III occur when the material in the overlay sample acts as a
'strain-relieving' layer. Mode II occurs when the material in the overlay
sample 'reinforces' the overlay. This can only occur if the material has a
higher modulus and sufficient cross-sectional area to substantially streng-
then the overlay. The same is true of geotextiles in the field as in samples in
the overlay tester. The two modes of behavior of the geotextiles may be
analyzed by use of fracture mechanics. In the 'strain-relieving' m o d e , a
crack normally begins at the crack in the old pavement, grows upward
quickly to the b o t t o m of the geotextile, and then stalls there for a while.
R e p e a t e d applications of traffic and thermal stresses initiate the crack
again on the upper side of the geotextile and impel its growth upward to
the surface of the overlay. In the 'reinforcing' mode, a crack in the old
pavement grows upward to the b o t t o m of the geotextile as before, but at
that point, it turns and propagates horizontally below the geotextile out to
a distance where insufficient energy is left to propagate the crack further.
Two analyses are n e e d e d in order to determine, in both the 'reinforcing'
m o d e and in the 'strain-relieving' mode, how much thickness of overlay
must be placed over the geotextile in order to achieve a desired reflection
cracking life. The essential elements of both design equations are given
in the following sections.

I I l-- Geotextile

(a)

[ ° l" Geotextile

(b)

[ t
l Ceotextile

Fig. 4. Modes of sample failure. (a) Failure Mode I: Crack propagates from bottom to top;
(b) Failure Mode II: Crack penetrates to the bottom of the geotextile and then develops a
slippage plane below the geotextile; (c) Failure Mode III: Crack propagates to the bottom
of the geotextile, then starts again at the top of the sample and propagates downward.
224 Robert L. Lytton

STRAIN RELIEVING MODE FRACTURE MECHANICS

Investigations at the Ohio State University 3-5 and Texas A & M


University 6-9 have applied fracture mechanics to predict the fatigue life of
an asphalt mixture. The results indicate that the rate of crack growth in an
asphalt concrete overlay follows the empirical power law developed by
Paris & Erdogan: m

dc
dN - A(AK)" (1)

where
AK = the stress intensity factor amplitude at the tip of the crack,
A, n = fracture properties of the material,
c = the crack length,
N = the number of load cycles.
Analysis of the crack growth in both the b e a m fatigue and overlay tests
provides values of the two fracture properties, A and n. Schapery 11
developed theoretical expressions of the fracture properties A and n in
terms of the creep, tensile and fracture energy properties of the material,
both using linear viscoelastic theory 1~ and the m o r e general J-integral
t h e o r y 12 which accounts for crack growth in nonlinear viscoelastic mate-
rials. Thus, A and n were shown to be true material properties, and were
also shown to be closely related to one another. It was found that for both
traffic fatigue and thermal fatigue:

2
n = -- (2)
m

where
m = the slope of the log creep compliance versus log time curve.
Further, it was found that for traffic-associated cracking, and typical
asphalt concrete mixes:

n = - 2.2 - 0.5 (logAF) (3)


and for thermal cracking, with the same asphalt concrete mix:

n = - 0-72 - 0.42 (IOgAT) (4)

Expressions for the n u m b e r of load cycles to reach failure are derived


Geotextiles for reinforcement of asphalt concrete 225

f r o m integrating eqn (1) for each m o d e of cracking propagation: bending,


shearing and thermal opening.

foNf dN = f a A (AK)"
dc (5)
co

where
Nf -- the number of load or thermal stress cycles to cause failure,
Co = the initial crack length which includes the thickness of the old pavement
surface layer,
d -- the total thickness of the asphalt ( = Co + do),
do = depth of the overlay.
For load-related stresses, there are two stages of crack growth as s h o w n
in Fig. 5, in which non-dimensionalized stress-intensity factors are shown.
W h e n the crack length reaches b e t w e e n 0.5 and 0.65 of the c o m b i n e d
d e p t h of the asphalt layers, d e p e n d i n g u p o n the degree of aggregate

1"8 /
AGGREGATEINTERLOCKRATIO /v

t
1.6 ~ LOW
O o/

HIGH •
1.2f,0 DUE TO SHEARING~,,
e / ~"~
&/; JA'

¢:

o
, , , ,°\ \ \ I 1 l
0 ~ ~2 ~3 ~4 ~5 0.6 ~7 O~ ~9 1~
c/d

Fig. 5. Non-dimensional bending and shearing stress intensity factors versus non-dimen-
sionalized crack length (c/d, where c and d represent the crack length and the combined
thickness of the existing pavement surface and overlay, respectively).
226 Robert L. Lytton

0"007

0.006

0-006

~004

F-
0-003

0"002

~001

l I I I [ I [ I I
0 0"I 0-2 0,3 0,4 0.5 0'6 0.7 0'8 0"9 I" 0

cld
Fig. 6. Non-dimensionalized thermal stress intensity factors versus non-dimensionalized
crack length (c/d, where c and d represent the crack length and the combined thickness of
the existing pavement surface and overlay, respectively).

interlock, there will be no further crack growth due to bending. From that
point on, all load-related crack growth is induced by shearing stresses. For
thermal stresses, the crack continues to grow up through the entire
asphaltic concrete thickness in proportion with the non-dimensionalized
thermal stress intensity factor shown in Fig. 6. The ordinates of the curves
in Figs 5 and 6 are multiplied by the following factors to determine the
stress intensity factors.
For bending:

qe 2
Kbo = #2 d3/------Tsin (6)

For shearing:

q
Kso = ~ [1 + e-t~(sin # l - cos#l)] (7)
'~pva

For thermal stresses:

E
I<,o - (1 - ~ ) v ~ s,~A T (8)
Geotextiles for reinforcement of asphalt concrete 227

where
q the tire inflation pressure,
l = the length of the tire footprint,
d = the combined thickness of the old pavement surface thickness and the
overlay,
E,V ~-" the elastic stiffness and Poisson's ratio of the overlay material,
Ot = the thermal coefficient of expansion of the underlying old cracked
pavement surface layer,
S the crack spacing in the old cracked pavement surface layer,
A T = the maximum change of temperature in the old pavement surface layer,

(3~3) 1/4
j3 = (9)

K = the coefficient of subgrade reaction of the layers below the old surface
layer which is determined by non-destructive testing.
D e v e l o p m e n t of the theory presented above is given in detail in Ref. 2.
The n u m b e r of load cycles to reach failure by each of these mechanisms
is proportional to:
(for bending)

1
Neo ~ - - (10)
AFK~,o

(for shearing)

1
Nfs ~ ~ (11)
A FK sno

(for thermal opening)

1
Nft cx _ _ (12)
AF K'~o

It was found in calibrating these to field data, that the observed n u m b e r


of cycles to failure is predicted very reliably with a linear combination of
these three numbers of load cycles to reach failure.
More details on the use of these equations are found in the report of the
F H W A project, in three volumes,2 and its implementation is in a c o m p u t e r
program on diskette which is available from the F H W A . The relations
given above show all of the mechanistic relations between the material
properties and the reflection cracking life of the overlay. For each fracture
228 Robert L. Lytton

mechanism, the relation between the log of the number of cycles to failure
and the material properties is:
logNf = - n log K - logA (13)
In its essence, the smaller the values of AF, AT and n, the longer the
overlay will last. The next question is: how do geotextiles affect the values
of AF, AT, and n?
The test data shown in Fig. 7 were measured on the overlay tester for a
variety of fabrics, n u m b e r e d 1 through 10, and compared with the results
of tests run on control samples with no fabrics. The fabrics represented on
this graph include some that have been used extensively and others which
are only in the developmental stage. In order for a fabric to do well in the

FRACTURE PROPERTIES OF FABRICS


TEST SERIES 4

CONTROL
o SINGLE PTS. 5
• DOUBLE PTS.

-3 n

I I I I I 0
-5.0 -4.0 -3'0 -2.0 -I.0 \o
log A

n " - 0.25 - 0-792 lOgl 0 A (Control samples)

Fig. 7. Fracture properties of overlays reinforced with fabrics.


Geotextiles f o r reinforcement o f asphalt concrete 229

'strain-relieving' mode, it must plot below and to the left of the line for the
control samples. Most fabrics, including those in commercial use, repre-
sented by No. 3, were able to show some improvement over the control
sample. In making a systematic study of the crack resistance of the samples
with the fabrics, fabrics Nos 2, 4, 1, 7 and 10 were ranked in decreasing
order. Fabrics Nos 1 and 2 were woven fabrics and No. 7 was a knitted
glass fabricl In general, it appears that the heavier fabrics with the greater
amount of excess tack coat above the optimum amount are generally more
fracture retardant.
The optimum tack coat for each of the fabrics that were tested was
determined using a standard laboratory procedure that was devised at
Texas A & M University. 13 The procedure dips a 75 mm × 375 mm strip
of fabric into AC-10 asphalt heated to 121°C and then places the strip
between sheets of newspaper and irons it with a heated iron and light
pressure. An AC-10 asphalt is an asphalt cement with a viscosity of about
1000 poises at 60°C. The viscosity grade is the viscosity at 60°C divided by
100. The strip is removed from between the newspaper and placed
between other sheets of newspaper and ironed again. The process is
repeated until no more asphalt is absorbed by the newspaper. At this
point, light can be seen through the fabric and all asphalt that has been
loosely held in capillary tension has been removed. The asphalt content of
the fabric strip is expressed as a percentage of the dry weight of the fabric,
and this is approximately the same asphalt content that will produce the
maximum shear strength of the fabric-asphalt interface. To this amount of
asphalt must be added an additional amount of asphalt to satisfy the
'surface hunger' of the asphalt on each side of the fabric. On newly mixed
asphaltic concrete in the laboratory, this surface hunger has been found to
be 0.36 l / m e .
It is possible to get too much tack coat so that the excess asphalt bleeds
through the overlay or so that the shear strength is too low to sustain
braking and cornering stresses.

R E I N F O R C I N G M O D E ANALYSIS

A simpler method of analysis is used for the reinforcing mode of geotextile


behavior. Those geotextiles which have worked well in this mode have
always had a larger elastic stiffness than the surrounding asphalt concrete.
Materials such as polypropylene grids and glass fabrics or grids are
appropriate here. It is useful at this point to define the terms used in
analyzing the overlay test data where the 'debonding' mode of failure
occurred.
230 Robert L. Lytton

I~ ~/2 _ I
fto~ Ido (..~-~ Stress =0
ff , .... . . . . . ~ . . . . .
To

(a)

~ f r o

Stress=0 - - - ~ ld u ~ S t r e s s =0
t------~- F
~'-~ ~ ~L'~BottomPlaten
(h)
~o

TO m n x

,, I rmin
UmGx U

(c)
Fig. 8. (a) Upper part of overlay sample; (b) lower part of overlay sample; (c) assumed
shear stress versus shear displacement relation.

The debonding analysis required writing the equations of horizontal


equilibrium of the upper and lower portion of the overlay sample, as
shown in Fig. 8. The upper portion includes both the fabric and the upper
layer of the overlay sample which has a depth, do, and a length, l/2, where 1
is the overall length of the sample. The lower layer of the overlay has a
depth, du, and a debonding length, lu. In order to open the crack to its full
width, A, a force F must be exerted on the bottom platen. The assumed
shear stress, z0, versus shear displacement, u, relation is shown in Fig. 8(c).
The relation has an initial slope, k0, a maximum shearing stress, ZOmax,and
a minimum final shear stress, Zmin.
There are two distinct stages in the debonding of a sample. Stage 1 is
before any debonding has occurred and Stage 2 is after it has been initiated
and the bedonding length, lu, as shown in Fig. 8(b), is greater than zero.
Geotextiles for reinforcement of asphalt concrete 231

Stage 1: Before debonding

In interpreting the following equations it is useful to refer to Figs 8(a)


(overlayer), 8(b) (underlayer), and 8(c) (shear stress versus shear dis-
placement relation).
The equation for the maximum force in the overlayer is: 14
F = b[ft0d0 +ff/f] (14)
where
b = the width of the sample,
ft0 = the tensile stress in the overlayer,
tf = the thickness of the fabric,
fftf = the tensile force per unit width in the fabric.
The equation for the same maximum force in the underlayer is

F = bA
(ko)(l)
-- l
/3 sinh/3-~-
cosh/3 ~ - 1
(15)

where

/3 =
•/ ko
Eu du
(16)

Eu = the elastic stiffness of the underlayers.

It is notable that if the strain in the overlayer is measured, then the stress
in the overlay, fto, can be found, and if the force, F, is measured, the force
per unit width in the fabric, fftf, can be determined from eqn (14).

Stage 2: After debonding

Three equations are found from the conditions for horizontal equilibrium
of the overlayer and the underlayer. The equation for the maximum force
in the overlayer after debonding has begun is identical with eqn (14).
F = b[ftodo +fftf] (17)
The equation for maximum force in the underlayer after debonding is:

F = b,rmialu+ A sinhfl~-sinh/31u
) +(B coshfl~
l cosh~lu
)]
(18)
232 Robert L. Lytton

where

A, B = constants that are evaluated from the boundary conditions of the


underlayer and will be defined later.
The crack opening, A, is given by eqn (19):

Eudu 2 ~-coshfllu ] +B[ sinhfl~-sinh/3lu


' ] (19)

The constants A and B are given by the following two equations:

rmi n l u 1
Eu d-------7-sinh/3
A = (20)
fl[sinh(fll~)sinh(fll) -cosh(fllu)cosh(fl~)]
rminlu l
Eo d----7cosh t3
B = (21)
[ sin,., ) - cos,.,,,,,u,COS
The equations given above contain several measured values, namely F,
A, do, du and l, and several unknowns, namely rmin, lu, k0, Eu, ft0, and fftf.
An analysis of the equations shows that it is impossible to solve for the
unknowns without making one more measurement or an assumption.
Because of this, the displacement in the overlayer, AI, was measured and
the equation relating the stress in the overlayer to this measured displace-
ment was:

~0 ~u(~) ,22,
Eu = the elastic stiffness of the underlayer which was assumed to equal that of
the overlayer.

DESIGN EQUATIONS: REINFORCING MODE

In designing an overlay using geotextiles in the reinforcing mode, a


selection must be made of the overlay thickness, do, the maximum force
that can be carried by the fabric or grid, the width of the fabric strip, and
the amount of tack coat used. Since the choice of one affects each of the
Geotextiles for reinforcement of asphalt concrete 233

others, it is possible, through an interactive process, to determine an


optimum overlay design.
For fabrics, the overlay thickness is:

do= k°A ~w m 1 ] - ntf


[ cosh--~- (23)
(ft0)/3 sinh(-~2~-) L

For grids, the overlay thickness is:


ko A
ao= I cosh - ~ - 1] (24)
(fto)(i+np) ¢1sinh (-0-~)

where all of the variables have been defined before except the following:
fro = the design tensile strength of the overlay,
w = the minimum width of fabric or grids placed above the old pavement
surface crack,
n = the elastic stiffness ratio, i.e. the ratio of the elastic stiffness of the grid or
fabric to the elastic stiffness of the overlay,
p = the ratio (in decimal form) of the cross-sectional area of the grid or fabric
divided by the cross-sectional area of the overlay.
A careful inspection of the design equations shows the following:
(1) The required thickness of overlay, do, increases as the crack open-
ing, A, increases.
(2) The required thickness of overlay decreases as the stiffness of the
fabric, ntf, increases.
(3) There is a complex interaction between the shear stiffness of the
tack coat, k0, the width of the fabric and the elastic stiffness and
thickness of the underlayer, Eu and du, that also affects the design
overlay thickness, do.
(4) For grids, the required thickness of overlay decreases as the cross-
sectional area ratio, p, and the elastic stiffness ratio, n, increase.
(5) When the calculated value of do is larger than the actual overlayer
thickness, fracture failure will occur. Conversely, when the calcu-
lated value is smaller than the actual thickness, debonding will
occur.
(6) If the fabric is placed continuously rather than in strips above each
crack in the old surface, then the fabric width, w, to use in the design
equations is the crack spacing. There is a slight increase of overlay
thickness as the crack spacing increases.
234 Robert L. Lytton

When placing strips of fabric above the cracks in an old pavement, the
minimum width of the strip can be determined by setting eqns (17) and
(18) equal. The minimum width of fabric strip is:

ft0
w = (do + ntf) (25)
Tmin

An inspection of this equation shows that the width of the strip i n c r e a s e s


with increasing overlay depth, fabric thickness and modulus, design tensile
strength of the overlay and with decreasing minimum shear strength of the
tack coat. Experimentation with the values of this minimum shear strength
has led to the conclusion that the viscosity and the film thickness of the tack
coat are determining factors in whether a 'reinforcing' mode of failure
occurs. Usually debonding can be assured if AC-10 or AC-20 is applied at
more than 0.68 l/m 2 above the optimum. A graph of minimum shear
strength versus the excess tack coat above optimum is given in Fig. 9. All
tests were made at room temperature (25°C). As is seen in the graph, the
minimum shear strength below which debonding occurs is usually less than
48 kN/m 2. The value of "rmin is important in determining the minimum
width of the fabric to prevent reflection cracking from the edges of the
fabric strip.
Values of the shear stiffness, k0, vary between 1250 and 2900 MN/m 3 for
AC-10 tack coat and between 2800 and 8000 MN/m 3 for AC-20 tack coat.
In general, the heavier fabrics produce the higher shear stiffness values.
All tests were made at normal room temperature (25°C).
The reinforcing mode of geotextile behavior can, if properly con-
structed, turn a reflection crack into a horizontal plane beneath the
reinforcing layer and thus delay the reflection cracking indefinitely. This is
the more desirable mode of geotextile behavior.

C A U T I O N A R Y R U L E S OF T H U M B

Geotextiles can be effective in retarding or delaying indefinitely the


appearance of reflection cracks, but these are only under certain restricted
conditions. The following have been found to be good rules of thumb
delineating the ranges of movement in the old pavement surface for which
geotextiles have been successful.
(1) S h e a r i n g . If the load efficiency factor across a crack is greater than
0-80, the geotextile has a better chance of working. The load
Geotextiles for reinforcement of asphalt concrete 235

I00 l-

90

80

(M
E 70
Z

-1-
)-
60

Z
b..l
re"
F- 50

-I- 40

~-
Z
30

I0
20

I0

0 L ~
0.0 0.25 0.50 0.75 1.00
EXCESS TACK COAT ABOVE OPTIMUM, Liters/sq.m

Fig. 9. M i n i m u m shear strength as a function of excess tack coat above optimum.

efficiency factor is the ratio of the deflection on the unloaded side


of a crack to the deflection on the loaded side. In general, geotex-
tiles cannot reinforce an overlay against shearing displacements.
(2) Thermal opening. There are three ranges of thermal opening: (a)
from 0.00 to 0.76 mm, when no geotextile is needed; (b) from 0-76
to 1-78 mm, which is the effective range of geotextiles; and (c)
greater than 1-78 mm, which is an opening movement that geotex-
tiles cannot normally withstand.
236 Robert L. Lytton

CONCLUSIONS

The analysis and design of asphaltic concrete overlays using geotextiles


may recognize two modes of behavior: 'strain relieving' for geotextiles
with an elastic stiffness lower than the surrounding asphalt concrete, and
'reinforcing' for those geotextiles with elastic stiffnesses larger than
asphaltic concrete, including polypropylene grids and glass fabrics and
grids. Equations are presented in this paper which lead to the selection of
the overlay thickness in conjunction with the properties of the geotextiles
both in the strain-relieving and reinforcing modes. Strain-relieving geotex-
tiles only retard reflection cracking; reinforcing geotextiles, properly
constructed, have a chance to re-direct the reflection crack and thus to
delay its appearance indefinitely. An important aspect of the 'reinforcing'
behavior is the quantity of tack coat used to attach the geotextile to the
layer below it. The tack coat must always be somewhat above the
'optimum' level as defined by the method for determining it that is outlined
in this paper, but not so great as to cause a significant loss of shear strength
on the horizontal plane below the geotextile. The slight excess of tack coat
is thought to assist in waterproofing the fabric from the infiltration of water
in the event that a crack reflects through to the surface.

REFERENCES

1. Lytton, R. L. & Kohutek, G. L., Evaluation of AASHO road test satellite


and environmental studies. Phase II Final Report, NCHRP Project 20-7,
Task 17, Texas Transportation Institute, 1983.
2. Jayawickrama, P. W., Smith, R. E., Lytton, R. L. & Tirado, M. R.,
Development of asphalt concrete overlay design equations. Final Report,
Volumes 1, 2 and 3, March, 1987.
3. Majidzadeh, K. C., Buranarom, C. & Karakouzian, M., Application of
fracture mechanics for improved design of bituminous concrete. Report No.
FHWA-RD-76-91, Vol. 1, June, 1976.
4. Majidzadeh, K., Ramsamooj, D. V. & Fletcher, T. A., Analysis of fatigue of
sand-asphalt mixtures. In Proceedings, Association of Asphalt Paving Tech-
nologists, Vol. 30, 1969.
5. Majidzadeh, K. & Ramsamooj, D. V., Development of testing procedures
and a method to predict fatigue failures of asphalt concrete paving systems.
The Ohio State University Research Foundation, Final Report, Project RF
2873, March, 1971.
6. Germann, F. P. & Lytton, R. L., Methodology for predicting reflection
cracking life of asphalt concrete overlays. Report No. FHWA/TX-79/
09 ÷ 207-5, March, 1979.
7. Lytton, R. L. & Shanmugham, U., Fabric reinforced overlays to retard
reflection cracking. Report RF-3424-5, Texas A&M Research Foundation,
February, 1983.
Geotextilesfor reinforcementof asphaltconcrete 237

8. Lytton, R. L. & Tseng, K. H., Laboratory evaluation of overlays reinforced


with fabrics. Texas Transportation Institute, May, 1983.
9. Lytton, R. L. & Jayawickrama, P. W., Reinforcing fiber glass for asphalt
overlays. Texas Transportation Institute, May, 1986.
10. Paris, P. C. & Erdogan, F., A critical analysis of crack propagation laws.
Transactions of the ASME, Journal of Basic Engineering, Series D, 85, No. 3,
1963.
11. Schapery, R. A., A theory of crack growth in viscoelastic media. Report
NN-2764-73-1, Mechanics and Materials Research Center, Texas A&M
University, 1973.
12. Schapery, R. A., Models for damage growth and fracture in nonlinear
viscoelastic particulate composites. In Proceedings, 9th US Congress of
Applied Mechanics, American Society of Mechanical Engineers, Book No.
H00228, 1982.
13. Pickett, D. L. & Lytton, R. L., Laboratory evaluation of selected fabrics for
reinforcement of asphalt concrete overlays. Report No. FHWA/TX-84/
20 + 271-1, Texas Transportation Institute, August, 1983.
14. Kennepohl, G. J. A. & Lytton, R. L., Pavement reinforcement with geogrids
for reflection cracking reduction. In Proceedings, Paving In Cold Areas,
Mini-Workshop, Canada~Japan Science and Technology Consultations,
Technical Memorandum of PWRI No. 2136, October 1984, pp. 863-82.

You might also like